vocal learning via social reinforcement by infant marmoset...

16
Report Vocal Learning via Social Reinforcement by Infant Marmoset Monkeys Highlights d Development of marmoset contact calls is influenced by contingent parental feedback d Use of twin infants controlled for genetics, perinatal experience, and growth d This is the first experimental evidence for vocal production learning in infant monkeys Authors Daniel Y. Takahashi, Diana A. Liao, Asif A. Ghazanfar Correspondence [email protected] (D.Y.T.), [email protected] (A.A.G.) In Brief Takahashi et al. show that infant marmoset monkeys are vocal learners. In brief but almost daily sessions, infant marmoset twins were experimentally provided with high or low levels of contingent parental vocal feedback to their vocalizations. More parental feedback accelerated the transition to consistently producing mature contact calls. Takahashi et al., 2017, Current Biology 27, 1844–1852 June 19, 2017 ª 2017 Elsevier Ltd. http://dx.doi.org/10.1016/j.cub.2017.05.004

Upload: others

Post on 24-Aug-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

Report

Vocal Learning via Social R

einforcement by InfantMarmoset Monkeys

Highlights

d Development of marmoset contact calls is influenced by

contingent parental feedback

d Use of twin infants controlled for genetics, perinatal

experience, and growth

d This is the first experimental evidence for vocal production

learning in infant monkeys

Takahashi et al., 2017, Current Biology 27, 1844–1852June 19, 2017 ª 2017 Elsevier Ltd.http://dx.doi.org/10.1016/j.cub.2017.05.004

Authors

Daniel Y. Takahashi, Diana A. Liao,

Asif A. Ghazanfar

[email protected] (D.Y.T.),[email protected] (A.A.G.)

In Brief

Takahashi et al. show that infant

marmoset monkeys are vocal learners. In

brief but almost daily sessions, infant

marmoset twins were experimentally

provided with high or low levels of

contingent parental vocal feedback to

their vocalizations. More parental

feedback accelerated the transition to

consistently producing mature contact

calls.

Page 2: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

Current Biology

Report

Vocal Learning via Social Reinforcementby Infant Marmoset MonkeysDaniel Y. Takahashi,1,2,* Diana A. Liao,1 and Asif A. Ghazanfar1,2,3,4,*1Princeton Neuroscience Institute2Department of Psychology3Department of Ecology and Evolutionary BiologyPrinceton University, Princeton, NJ 08544, USA4Lead Contact

*Correspondence: [email protected] (D.Y.T.), [email protected] (A.A.G.)

http://dx.doi.org/10.1016/j.cub.2017.05.004

SUMMARY

For over half a century now, primate vocalizationshave been thought to undergo little or no experi-ence-dependent acoustic changes during develop-ment [1]. If any changes are apparent, then they areroutinely (and quite reasonably) attributed to thepassive consequences of growth. Indeed, previousexperiments on squirrel monkeys and macaquemonkeys showed that social isolation [2, 3], deafness[2], cross-fostering [4] and parental absence [5] havelittle or no effect on vocal development. Here, weexplicitly test in marmoset monkeys—a very vocaland cooperatively breeding species [6]—whetherthe transformation of immature into mature contactcalls by infants is influenced by contingent parentalvocal feedback. Using a closed-loop design, weexperimentally providedmore versus less contingentvocal feedback to twin infant marmoset monkeysover their first 2 months of life, the interval duringwhich their contact calls transform from noisy, imma-ture calls to tonal adult-like ‘‘phee’’ calls [7, 8]. Infantswho received more contingent feedback had a fasterrate of vocal development, producingmature-sound-ing contact calls earlier than the other twin. The dif-ferential rate of vocal development was not linkedto genetics, perinatal experience, or body growth;nor did the amount of contingency influence theoverall rate of spontaneous vocal production. Thus,we provide the first experimental evidence for pro-duction-related vocal learning during the develop-ment of a nonhuman primate.

RESULTS

Marmoset monkeys (Callithrix jacchus) are a voluble New World

species that exhibit a complex system of vocal communica-

tion [9]. This system includes vocal turn-taking: two adult

marmosets (related or unrelated to each other and out of sight

from one another) will exchange extended, phase-locked se-

quences of contingent phee calls, a type of contact call, acting

1844 Current Biology 27, 1844–1852, June 19, 2017 ª 2017 Elsevier

in essence like coupled oscillators [10]. Developmentally, infant

marmosetmonkeys produce long bouts of vocalizations consist-

ing of both immature- and mature-sounding calls [7, 8, 11, 12].

A subset of these calls are immature versions of the phee call

[7, 8], and the timing at which these immature calls transform

into mature-sounding calls varies widely among infants [7].

One hypothesis to account for this developmental variation is

differential feedback from parents. In this scenario, the contin-

gent contact call exchanges that adults use with each other

could also be used by parents as they respond to infant calls.

This could act as a ratchet for contact call development. For

example, studies of naturalistic human infant-parent interactions

[13–16], as well as experimental studies [17, 18], reveal that

contingent parental responses accelerate the development of

infant vocalizations, making them sound more mature (i.e.,

speech-like). Thus, perhaps marmoset parents that produce

contingent vocal responses to infant vocalizations similarly

accelerate their development. Indeed, in marmoset monkeys,

there is a strong correlation between the amount of contingent

vocal feedback from parents and the maturation rate of contact

calls [7]. There is no correlation, however, between the overall

amount of exposure to parental vocalizations and vocal develop-

ment [7]. This has led to the suggestion that developing

marmoset monkeys—unlike every other nonhuman primate

investigated thus far—may be vocal learners [19]. A viable alter-

native hypothesis is that, instead of an instance of vocal learning,

marmoset parents are simply responding more to healthier in-

fants who develop their vocalizations more quickly than others.

We designed an experiment to explicitly test whether or not

contingent vocal feedback can increase the rate at which

marmoset infants begin producing mature-sounding contact

calls. Because marmoset monkeys typically give birth to dizy-

gotic twins [20], we could control for the influence of genetics

and the perinatal environment on vocal development [8]. We

tested three pairs of twins (six infants) from three different sets

of parents. Starting at postnatal day 1 (P1), infants were briefly

removed from their home cage and provided different levels of

contingent feedback using closed-loop, computer-driven play-

backs of parental phee calls (Figures 1A and 1B). One randomly

selected twin was provided the best-possible simulated

‘‘parent’’ who provided 100% vocal feedback at an �1 s delay

if the infant produced a low-entropy contact call, i.e., a more

mature-sounding call; the other infant was provided a not-so-

good ‘‘parent’’ and received vocal feedback for only 10% of

Ltd.

Page 3: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

325cm

290c

m 0

10

20

0 5 10 15 20 25 30 350

10

20

Time (s)

Freq

uenc

y (k

Hz)

High playback response rate = ~100%

Low playback response rate = ~10%

occluder = Playback

A B

Test (no stim.)

Playback

30 min5-10 min

C

2

6

10

14

18

-40

-20

0

-2

0

2

4

-60

2 6 10 14 18 1ms

Higher AM frequencyHigher dominant frequency

Higher Wiener entropy = morebroadband signal

Freq

uenc

y (k

Hz)

Pow

er (d

B)

Am

plitu

de (n

u)

Time (s) Frequency (kHz)0.2 0.6 1.0

Shorter duration

0.2 0.6

-60

-40

-20

0

-1

0

1

2

6

10

14

18

1.0 2 6 10 14 18 1ms

Lower AM frequencyLower dominant frequency

Lower Wiener entropy = morenarrow band signal

Freq

uenc

y (k

Hz)

Pow

er (d

B)

Am

plitu

de (n

u)

Longer duration

Vocal folds

Upper vocal tract

Lungs

D

Wiener entropy

AM frequency

Duration

Dominant frequency

Upper vocal tract shape Vocal fold tension RespirationChanges in acoustics are consequence of changes in

Time (s) Frequency (kHz)

Younger infant

Older infant

Figure 1. Experimental Design and Potential Acoustic Parameters

(A) Infants were briefly separated from their parents and placed in an acoustically treated testing room. Computer-controlled playbacks were delivered through a

speaker. Sessions lasted �40 min, with the first 5 min (postnatal days 1 to 7) or 10 min used to collect spontaneous vocalizations.

(legend continued on next page)

Current Biology 27, 1844–1852, June 19, 2017 1845

Page 4: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

the low-entropy contact calls it produced (Figure 1B). The use of

an �1 s delay is based on data collected under naturalistic con-

ditions showing that parents usually provide vocal responses at

around 1 s after an infant vocalization (Figure S1) [21]. The high-

versus low-contingent response rates are respectively higher

and lower than the average parental contingent response rate

during naturalistic infant-parent interactions (21.35% ± 0.17%,

mean ± SEM) [7]. Each experimental session lasted 40 min; the

first 10 min (5 min in the first week, when the infants were neo-

nates) was used to record the infants’ spontaneous vocaliza-

tions, and the remaining 30 min was used for playback; the in-

fants were otherwise with their families for the remaining 23 hr

and 20min of each day. The use ofmaternal versus paternal con-

tact calls was counter-balanced on a session-by-session basis.

We used 20 call exemplars from each parent. These conditioning

sessions occurred almost every day for 2 months (�14 consec-

utive experiment days + 1 rest day, iterated four times).

We used multiple acoustic measures to quantify develop-

mental changes as a continuous process [7, 22] (Figure 1C).

We did this for two reasons. First, it allowed us to measure

change without the bias of ethological classifications. Second,

it allowed us to see whether some acoustic parameters versus

others were more sensitive to contingent parental vocal feed-

back. The different acoustic parameters can be related to

different biological mechanisms or their combination (Figure 1D).

For example, an optimal control-based Waddington landscape

model of marmoset vocal development revealed that changes

in dominant frequency could be completely accounted for by

the growth of the vocal tract [23]. The four acoustic measures

that we used were duration, dominant frequency, amplitude

modulation (AM) frequency, and Wiener entropy, all of which

change over the course of development (Figure 1C) [7, 22]: dura-

tion increases, dominant frequency decreases, AM frequency

decreases, and Wiener entropy decreases as the infant marmo-

sets get older [7].

Because marmoset infants will immediately change (on the

order of seconds) their call acoustics to be more mature upon

hearing a contingent parental response [21], our acoustic mea-

sures were only performed on the spontaneous calls produced

by infants during the 5 or 10min interval at the beginning of every

session (Figure 1A). We used multiple linear regression to fit the

data with the relevant acoustic measure as the dependent vari-

able and with postnatal day (n = 193 days, six infants, 30–

33 days/infant), contingency group (high versus low), and twin

identity (1, 2, or 3) as the predictors. The inclusion of twin identity

allowed us to control for the effect of genetics in the develop-

ment. All one-way and two-way interactions were included to

account for relevant effects. All p values reported below are

calculated from the test of nullity of the interaction between post-

(B) Twin infants received either high-contingency playbacks (100%) or low-cont

were delivered relative to the infant vocalizations. Warmer colors indicate higher

(C) Four acoustic parameters change over the course of marmoset vocal develop

entropy, and amplitude modulation (AM) frequency. Vertical red dashed lines in s

panels.

(D) Four acoustic parameters are related to different operations of the vocal appa

frequency changes are associated with changes in the shape of upper vocal trac

associated with size of upper vocal tract. Change in duration is associated with

See also Figure S1.

1846 Current Biology 27, 1844–1852, June 19, 2017

natal day and contingency group; we set our alpha level at 0.01.

We also report the adjusted R2 (adj. R2) of the regression

model. The coefficients of the main regression models are re-

ported in the Supplementary Information. Because we wanted

to capture the rate of vocal change up until the point at which

the infant marmosets produce only mature-sounding contact

calls (>95% phee calls), the regression analyses were done on

ages P1 to P35. For all infants (n = 6), mature-sounding phee

calls were produced almost exclusively after P36, and no effect

of contingency group was observed for any of the four acoustic

measures after P36 (Figure S2).

We present the Wiener entropy data first, because this mea-

sure effectively captures the transformation of noisy, immature

(high-entropy) contact calls into tonal, adult-like (low-entropy)

calls [7, 8, 21]. For each twin pair, the individual that received

more contingent feedback had a faster rate of vocal develop-

ment, producing mature-sounding (lower-entropy) calls earlier

than the other twin. Figure 2A shows that the timing of the tran-

sition from immature to mature calls was quicker for the infants

that received more contingent feedback (adj. R2 = 0.519,

p = 0.0022). Figure 2B shows that this pattern held true for

each pair: the individual that received more contingent feedback

had a steeper rate of vocal development, producing lower-

entropy contact calls more quickly than the other twin. Measures

of AM frequency revealed a similar pattern. Figure 2C shows

that the development of this acoustic parameter was also faster

in infants receiving high- versus low-contingent vocal feedback

(adj. R2 = 0.490, p = 0.0068). Again, this pattern held true for

each pair, whereby the individual that received more contingent

feedback developed the mature AM frequency more quickly

when compared to the other twin (Figure 2D). It is possible that

part of the differences in the Wiener entropy and AM frequency

development are due to initial differences in vocal behavior

exhibited by the infants immediately after birth. To verify this

possibility, we tested whether the intercepts of the regression

models were different between contingency groups and found

no evidence for this (test of nullity for the mean effect of

contingency group, Wiener entropy: p = 0.7851; AM frequency:

p = 0.0715).

Because the twins were not identical in their growth rates (Fig-

ure 2E), one possibility is that growth accounts for vocal develop-

mental differences. Body weight is a good proxy for overall

growth, and weight correlates well with vocal apparatus size in

monkeys [24]. We therefore added body weight and its interac-

tion with postnatal day as predictors. The result revealed that dif-

ferences in weight cannot explain the differential development of

the Wiener entropy or AM frequency changes as a function of

high versus low contingency (Wiener entropy: adj. R2 = 0.494,

p = 0.0002; AM frequency: adj. R2 = 0.568, p = 0.0056).

ingency playbacks (10%). Spectrograms depict when such playbacks (green)

values.

ment and were measured in the study: duration, dominant frequency, Wiener

pectrograms indicate the time interval used for the analyses in the neighboring

ratus (vocal tract, vocal folds, and lungs/respiration). Wiener entropy and AM

t, vocal fold tension, and respiratory control. Change in dominant frequency is

change in lung capacity and respiratory control.

Page 5: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

-28

-24

-20

-16

5 10 15 20 25 30Postnatal day

Wie

ner e

ntro

py (d

B)

A

700

800

900

1000

1100

-40

-20

0

-40

-20

0

-40

-20

0

600

1000

1400

1800

300

500

700

900

500

1500

355 10 15 20 25 30Postnatal day

35

AM

freq

uenc

y (H

z)

10 20 30 10 20 30 10 20 30

Ent

ropy

(dB

)

10 20 30 10 20 30 10 20 30

AM

freq

. (H

z)

B

C

D

High contingentLow contingent

n = 193, p = 0.0068, Adj R2 = 0.490n = 193, p = 0.0022, Adj R2 = 0.519

High contingentLow contingent

Twin set 1 Twin set 2 Twin set 3 Twin set 1 Twin set 2 Twin set 3

10 20 30

30

40

50

10 20 30

40

50

60

70

80

10 20 30

30

40

50

60

E

Wei

ght (

g)

Twin set 1 Twin set 2 Twin set 3

High contingentLow contingent

Postnatal day

Postnatal day Postnatal day

Figure 2. Infants Receiving More Contingent Vocal Feedback Develop Their Vocalizations Faster, and This Change Is Not Related to Growth

Differences

(A and B)Wiener entropy (in decibels) changes over postnatal days for high- and low-contingency infants. (A) shows group average; shaded regions indicate 1 SE

intervals. (B) shows data for each twin set.

(C and D) AM frequency (in Hz) changes over postnatal days for high- and low-contingency infants. (C) shows group average; (D) shows data for each twin set.

(E) Growth of all infants as measured by weight (in g) over postnatal days.

See also Figures S2 and S3.

In contrast to Wiener entropy and AM frequency, Figures 3A

and 3C show that developmental changes in call duration and

dominant frequency were not influenced by the amount of

contingent feedback (duration: adj. R2 = 0.507, p = 0.0527; domi-

nant frequency: adj. R2 = 0.557, p = 0.6278). We also didn’t find

differences in the intercepts between contingency groups (dura-

tion: p = 0.2214; dominant frequency: p = 0.0625). These pat-

terns held true for each twin set (Figures 3B and 3D). The same

result holds when body weight and its interaction with postnatal

day are included as predictors in the regression (duration: adj.

R2 = 0.469, p = 0.2697; dominant frequency: adj. R2 = 0.546,

p = 0.3008).

Because the different contingency conditions necessarily

result in one infant being provided with more parental calls

than the other (Figure S3), another possibility is that the total

amount of playback calls—and not their contingency—can

explain the learning effect on the rate of call development.

This possibility is not supported by our previous study, which

showed that there was a strong correlation between the timing

of contact call development and contingency but no correlation

with the much greater number of non-contingent parental calls

heard by the infant [7]. Nevertheless, we directly tested this hy-

pothesis by including the total number of playback call re-

sponses and its interaction with postnatal day as predictors in

the regression. The regression result shows that even after

controlling for the total number of playback stimuli, the contin-

gency-driven vocal developmental changes remain statistically

significant (Wiener entropy: adj. R2 = 0.506, p = 0.0038; AM

Current Biology 27, 1844–1852, June 19, 2017 1847

Page 6: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

9.2

9.6

10.0

0.3

0.4

0.5

0.6

0.1

0.3

0.5

0.7

0.2

0.4

0.6

0.8

0.2

0.4

0.6

0.8

9

10

11

8

9

10

11

9

10

11

5 10 15 20 25 30Postnatal day

35 5 10 15 20 25 30Postnatal day

35

Dur

atio

n (s

)

Dom

inan

t fre

quen

cy (k

Hz)

A

10 20 30 10 20 30 10 20 30

Dur

atio

n (s

)

10 20 30 10 20 30 10 20 30D

om fr

eq. (

kHz)

B

C

D

n = 193, p = 0.0527, Adj R2 = 0.507 n = 193, p = 0.6278, Adj R2 = 0.557

High contingentLow contingent

High contingentLow contingent

Twin set 1 Twin set 2 Twin set 3 Twin set 1 Twin set 2 Twin set 3

Postnatal day Postnatal day

Figure 3. Call Duration and Dominant Frequency Are Not Influenced by the Amount of Vocal Feedback

(A and B) Duration (in s) changes over postnatal days for high- and low-contingency infants. (A) shows group average; shaded regions indicate 1 SE intervals.

(B) shows data for each twin set.

(C and D) Dominant frequency (in kHz) changes over postnatal days for high- and low-contingency infants. (C) shows group average; (D) shows data for each

twin set.

See also Figures S2 and S3.

frequency: adj. R2 = 0.509, p = 0.0040), while the other acoustic

parameters remain uninfluenced by the differential feedback

(duration: adj. R2 = 0.513, p = 0.0505; dominant frequency: adj.

R2 = 0.544, p = 0.3293).

Given that the difference in the amount of contingency feed-

back resulted in differential rates of vocal development, we

wondered if infants with a faster rate produce more calls in

general. A change in the rate of call production could indicate

a long-term change in the animal’s baseline state, e.g., stress

levels [25]. To test this, we fitted a multiple linear regression in

which the rate of infant call production (number of calls per min-

ute for each baseline period) was the dependent variable and the

predictors were the postnatal day, contingency group, twin set,

and all one- and two-way interactions. There was no significant

effect of contingency group (n = 193, adj. R2 = 0.538, p = 0.2306)

and no significant interaction between contingency group and

postnatal day (p = 0.6664) in the first 35 days of postnatal life

(Figure 4A). The same pattern was true later in the second month

of life, when infants are producing phee calls exclusively (post-

natal days 36 to 60; n = 130, adj. R2 = 0.145, contingency group,

p = 0.0919; interaction between contingency group and post-

natal day, p = 0.7599) (Figure 4B). Thus, contingent feedback

from parents does not influence the overall rate of vocal produc-

tion and cannot be linked to long-term differences in the state of

the animal.

Our sample size was small for each contingency group (n = 3).

Thus, it is important to estimate the statistical power of our re-

sults as a measure of their robustness. If the power is low, it

1848 Current Biology 27, 1844–1852, June 19, 2017

can lead to (1) a low probability of finding a true effect, (2) an

overestimation of the effect size when the true effect is discov-

ered, and (3) a low positive predictive value (PPV) when an effect

is claimed. Naturally, one major source of low power is a small

sample size, but sources of high statistical power include exper-

imental designs that control for different confounds like genetics,

stimuli, and context (as in the current study) [26]. We calculated

the effect size (Cohen’s f2), performed a power analysis, and esti-

mated the PPV of our results. For the regressionmodel with PND,

Group, TwinId, and their two-way interactions, the Cohen’s f2

values (with 90% CI) were 0.1841 ([0.0871, 0.3002]), 0.0115

([0, 0.0365]), 0.1093 ([0.0357, 0.1942]), and 0.0400 ([0, 0.0883]),

respectively, for Wiener entropy, AM frequency, duration, and

dominant frequency. For a = 0.01 (type I error), these effect sizes

imply a post hoc achieved power of 0.9985 (Wiener entropy),

0.0929 (AM frequency), 0.9491 (duration), and 0.4546 (dominant

frequency).

We can now estimate the positive prediction value as PPV =

([1 � b] 3 R)/([1 � b] 3 R + a), where (1 � b) is the power, a is

the type I error control, and R is the odds that a claimed effect

is indeed non-null among the effects being claimed [26]. Calcu-

lating the pre-study odds R is usually difficult, but we can get an

estimate using the literature. Among perturbation studies of non-

human primate vocal learning, one study observed a strong

effect of social contact [27], and three studies did not [2–4].

Therefore, we can assume a pre-study odds equal to 0.25 and

conclude that the PPV for each acoustic parameter is 0.9615

(Wiener entropy), 0.6990 (AM frequency), 0.9596 (duration),

Page 7: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

0

10

20

0

4

8

0

10

20

A High contingentLow contingent

40 50 60Postnatal dayPostnatal dayPostnatal day

Rat

e (c

alls

/min

)R

ate

(cal

ls/m

in)

B High contingentLow contingent

Twin set 1 Twin set 2 Twin set 3

Twin set 1 Twin set 2 Twin set 3

40 50 60 40 50 60

0

10

20

30

0 10 20 30Postnatal day

0 10 20 30Postnatal day

0 10 20 30Postnatal day

0

10

20

0

20

40

Figure 4. Development of the Rate of Call

Production Measured during the Baseline

Interval

(A) Rate of call production from postnatal day 1 to

postnatal day 35 for each infant. Orange and black

lines are cubic spline curves fitted to high- and

low-contingency data, respectively.

(B) Rate of call production from postnatal day 36

to postnatal day 60 for each infant. Orange and

black lines are cubic spline curves fitted to high-

and low-contingency data, respectively.

and 0.9191 (dominant frequency). We conclude that (1) the pos-

itive finding for Wiener entropy is highly robust, whereas the

finding for AM frequency is less robust; (2) there is not sufficient

evidence for the effect of contingency group on duration, but

there might be a smaller effect; and (3) there is no evidence for

the effect of contingency group on dominant frequency.

Overall, our finding that contingent feedback changes the rate

of infant vocal development is statistically robust and is consis-

tent with our previous correlation-based evidence [7].

DISCUSSION

It is generally accepted that there are three varieties of vocal

learning: comprehension, usage, and production [28]. Compre-

hension learning is when an animal learns to respond appropri-

ately—via experience—to vocal signals. For example, infant

vervet monkeys learn adaptive responses to alarm calls by

watching what adult conspecifics do [29], infant macaques learn

via experience to recognize their mother’s voice [30], and Diana

monkeys learn to respond adaptively to the alarm calls of other

species [31]. Usage learning is when an animal learns in which

context(s) to produce a pre-existing call in its repertoire. For

instance, infant and juvenile vervet monkeys produce adult-like

‘‘raptor’’ alarm calls but do so to the wrong birds; they eventually

learn to associate their alarm calling to the bird species [29].

Similarly, infant marmoset monkeys produce some call types in

inappropriate contexts [32], and learning the appropriate context

is experience dependent [27]. Another instance of usage learning

is the vocal conditioning of captive primates: macaques, for

example, can be trained to associate their vocalizations with

food or context cues [33, 34] and even to arbitrary visual cues

[35]. The third variety of vocal learning—production learning—

is the experience-dependent change in the acoustic structure

of vocalizations [28]. Increasingly, however, the literature has

limited the definition of vocal production learning to learning

novel vocalizations via imitation (e.g., [36]). Indeed, some have

Current B

limited it even further to include only

imitation of those vocalizations that

involve changes in the laryngeal/syrin-

geal control [37].

Our data show that infant marmoset

monkeysexhibit vocal production learning

via social reinforcement from parents.

They do this not through imitation, but

rather through the experience-depen-

dent increase in the control of the vocal

apparatus that allows them to more consistently produce tonal

(low-entropy) phee calls. Early in life, infant marmosets produce

cries, phee-cries, and subharmonic phees that are immature ver-

sions of the phee call. Relative to the phee call (which they also

produce, but inconsistently), these immature versions are higher

in spectral entropy, dominant frequency, and amplitudemodula-

tion frequency, and shorter in duration [7, 8, 21, 23] (Figure 1C).

The production of these immature contact calls is related to

poor muscular control of—and coordination between—respira-

tion and vocal fold tension [7, 8, 23]. These immature phee calls

disappear later in development; they are not produced in any

other contexts. The developmental timing of this control and

coordination of vocal apparatus elements is what is linked to

experience: more contingent vocal feedback results in faster

development of respiratory and laryngeal control and coordina-

tion [23]. These data support the notion that forms of vocal pro-

duction learning extend beyond imitation [38, 39].

Marmoset monkeys prove to be exceptional nonhuman pri-

mateswith regard to vocal learning. Not only do theyundergodra-

matic changes in the acoustic structure of their vocalizations dur-

ing development [7, 8], but two recent studies also showed the

possibilityof vocal learning via experiencewithparents in this spe-

cies. In the first study, there was a strong correlation between

contingent parental feedback and the rate at which immature ver-

sions of the contact phee call transformed intomature versions of

the call [7]. In the second study, infant marmoset triplets with

limited parental contact (one infant was hand-raised and other

two were separated after 3 months) had their vocal behavior as

sub-adults compared to their normally reared twin siblings from

another pregnancy [27]. The vocal behavior of the triplets with

limited parental contact remained immature relative to their nor-

mally reared twin siblings by multiple measures, demonstrating

that some aspect of parental care is essential for normal vocal

development in marmoset monkeys [27]. The present study

experimentally demonstrated the specific role of parental vocal

feedback in influencing vocal production development. By using

iology 27, 1844–1852, June 19, 2017 1849

Page 8: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

twins, we could reduce the potential confounds of genetics and

perinatal experience [8]. Moreover, our analyses showed that

neither growth nor the number of calls provided during the play-

back sessions influenced the rate of learning. The overall sponta-

neous call rate of each infant was not influenced by contingency,

either. This shows that social reinforcement is acting on a vocal

motor learning circuit and not simply changing baseline levels of

a behavioral state (e.g., arousal or stress levels).

The social reinforcement-based vocal learning by infant

marmoset monkeys is consistent with findings in experimental

studies of early (prelinguistic) vocal development in humans

[17, 18] and song learning in birds (e.g., cowbirds [40] and zebra

finches [41]). In human infants, for instance, immature vocaliza-

tions (e.g., cries and cooing sounds, among many others)

gradually transform into the consistent, context-dependent pro-

duction of speech-like babbling output [42, 43]. The speed of this

transformation is influenced by contingent parental feedback

[13–18]. Another similar development pattern occurs during

locomotion learning in humans. Human toddlers alternate be-

tween crawling and walking, and only later in development will

start to walk consistently with a more adult-like gait [44, 45].

This locomotor transition is also influenced by parental social

feedback [46]. Similarly, we found that marmoset infants that

received greater contingent vocal feedback than their twin

began to consistently produce mature-sounding phee calls

earlier in life. Our experimental design used a brief experimental

period (35–40 min); for the remaining 23+ hr of the day, the in-

fants were back with their families. Yet this brief experimental

manipulation period had a significant effect on infant vocal

development. This suggests that the cumulative effect of a

socially relevant learning cue (contingency) and repetition can

strongly influence the rate of vocal learning.

Given that brief but consistent socially relevant contingent

feedback can have a large impact on shaping infant vocal devel-

opment in humans [17, 18], birds [40, 41], and marmosets (cur-

rent study), do these animals share a similar learning substrate?

A shared neural substrate may involve links between the limbic

system and cortico-striatal systems [47]. These two systems

are common across vertebrate species. Contingent social re-

sponses drive the learning via the cortico-striatal system, and

the limbic system (which subsumes the evolutionarily conserved

social behavior network [48–50]) drives the motivation to attend

to social signals. Vocal learning by infant marmoset monkeys is

entirely consistent with this ‘‘social motivation-vocal learning’’

framework [47]. Developing marmoset monkeys also exhibit a

pattern of FoxP2 expression in their cortico-striatal circuit [51]

that is analogous to that of songbirds and humans [52], lending

credence to the notion that marmoset infants use this circuit to

guide their phee call development through social reinforcement,

just as birds and humans do. The fact that changes induced by

social reinforcement inmarmoset infants did not affect all acous-

tic parameters equally should help investigations of the underly-

ing vocal learning mechanism. The high-contingency condition

sped up changes in Wiener entropy and AM frequency but did

not influence changes in duration and dominant frequency. In

marmoset monkeys, changes in dominant frequency occur sim-

ply by a growth-related increase in the length of the vocal tract

[23]. Similarly, changes in contact call duration are likely the

result of a growth-related increase in lung capacity. By contrast,

1850 Current Biology 27, 1844–1852, June 19, 2017

our systems-level vocal development landscape framework re-

vealed that the production of tonal versus noisy contact calls

(as measured by Wiener entropy) is linked to a learning-related

increase in musculature control of the larynx and respiration,

as well as the coordination of that control across these two ele-

ments of the vocal apparatus [23].

Why do humans and marmoset monkeys exhibit similar

patterns of vocal development in the early postnatal period?

Because some 40million years have passed since the OldWorld

and NewWorld primate lineages split [53], we suggest that vocal

learning arose as a by-product of the convergent evolution of a

cooperative breeding system. Cooperative breeding is only

found in about 3% of mammals [54]. Of those mammals,

marmoset monkeys and others in the callitrichid family are

the only nonhuman primates known to exhibit this strategy

[54, 55]. For marmosets, the rearing of infants is greatly reliant

on a concerted effort among the breeding female, breeding

male, non-breeding siblings, and other familiar but unrelated

group members. Marmoset caregivers compete to carry infants

[56, 57] and frequently provision food to offspring [58, 59]. This

cooperative breeding framework, in which non-parents within a

social group spontaneously care for offspring other than their

own, drives a more general tendency to be prosocial in other

domains [60], including vocal communication [6].

How does this lead to vocal production learning in infant mar-

mosets and humans? Care of infants is probably themost impor-

tant context in which cooperation with unrelated individuals

occurs. There is a strong correlation between the amount of in-

fant care provided by others and the reproductive success of a

mother [61]. In an environment where caregiver attention is a

limited resource and where non-maternal caregivers may have

a higher threshold than mothers to provide care, evolution may

select for vocal behaviors that help infants attract caregiver

attention [62]. A related hypothesis is that human infant vocaliza-

tions that soundmore speech-like evolved to exploit pre-existing

auditory predispositions in adult receivers [63]. The fact that par-

ents of both human and marmoset infants are more likely to give

contingent responses to infant vocalizations when those vocali-

zations sound more adult-like [21, 64] is consistent with this

‘‘receiver predisposition’’ idea [63]. We therefore suggest that

the vocal learning mechanism evolved to speed up the produc-

tion of mature-sounding vocalizations (those that exploit the

receiver predispositions) using social feedback because such

vocalizations are more likely to elicit caregiver attention.

STAR+METHODS

Detailed methods are provided in the online version of this paper

and include the following:

d KEY RESOURCES TABLE

d CONTACT FOR REAGENT AND RESOURCE SHARING

d EXPERIMENTAL MODEL AND SUBJECT DETAILS

B Marmosets

d METHOD DETAILS

B Experimental setup

B Closed loop playbacks

B Detection of calls and quantification of acoustic pa-

rameters

Page 9: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

d QUANTIFICATION AND STATISTICAL ANALYSIS

B Multiple linear regression analysis

B Linear regression models

B Effect size and power analysis

d DATA AND SOFTWARE AVAILABILITY

SUPPLEMENTAL INFORMATION

Supplemental Information includes three figures and can be found with this

article online at http://dx.doi.org/10.1016/j.cub.2017.05.004.

AUTHOR CONTRIBUTIONS

D.Y.T.: conceptualization, data collection and curation, data analysis, writing

original draft, review and editing. D.A.L.: data analysis, draft review and edit-

ing. A.A.G.: conceptualization, supervision, writing original draft, review and

editing, funding acquisition.

ACKNOWLEDGMENTS

This work was supported by a Scholar Award from the James S. McDonnell

Foundation (A.A.G.) and a National Science Foundation Graduate Research

Fellowship (D.A.L.).

Received: February 25, 2017

Revised: April 3, 2017

Accepted: May 3, 2017

Published: May 25, 2017

REFERENCES

1. Egnor, S.E.R., and Hauser, M.D. (2004). A paradox in the evolution of pri-

mate vocal learning. Trends Neurosci. 27, 649–654.

2. Hammerschmidt, K., Freudenstein, T., and Jurgens, U. (2001). Vocal

development in squirrel monkeys. Behaviour 138, 1179–1204.

3. Winter, P., Handley, P., Ploog, D., and Schott, D. (1973). Ontogeny of

squirrel monkey calls under normal conditions and under acoustic isola-

tion. Behaviour 47, 230–239.

4. Owren, M.J., Dieter, J.A., Seyfarth, R.M., and Cheney, D.L. (1992). ‘Food’

calls produced by adult female Rhesus (Macaca mulatta) and Japanese

(M. fuscata) macaques, their normally-raised offspring, and offspring

cross-fostered between species. Behaviour 120, 218–231.

5. Hammerschmidt, K., Newman, J.D., Champoux, M., and Suomi, S.J.

(2000). Changes in rhesusmacaque ‘coo’ vocalizations during early devel-

opment. Ethology 106, 873–886.

6. Borjon, J.I., and Ghazanfar, A.A. (2014). Convergent evolution of vocal

cooperation without convergent evolution of brain size. Brain Behav.

Evol. 84, 93–102.

7. Takahashi, D.Y., Fenley, A.R., Teramoto, Y., Narayanan, D.Z., Borjon, J.I.,

Holmes, P., and Ghazanfar, A.A. (2015). The developmental dynamics of

marmoset monkey vocal production. Science 349, 734–738.

8. Zhang, Y.S., and Ghazanfar, A.A. (2016). Perinatally influenced autonomic

nervous system fluctuations drive infant vocal sequences. Curr. Biol. 26,

1249–1260.

9. Bezerra, B.M., and Souto, A. (2008). Structure and usage of the vocal

repertoire of Callithrix jacchus. Int. J. Primatol. 29, 671–701.

10. Takahashi, D.Y., Narayanan, D.Z., and Ghazanfar, A.A. (2013). Coupled

oscillator dynamics of vocal turn-taking in monkeys. Curr. Biol. 23,

2162–2168.

11. Elowson, A.M., Snowdon, C.T., and Lazaro-Perea, C. (1998). Infant ‘bab-

bling’ in a non-human primate: complex vocal sequences with repeated

call types. Behaviour 135, 643–664.

12. Pistorio, A.L., Vintch, B., and Wang, X. (2006). Acoustic analysis of vocal

development in a New World primate, the common marmoset (Callithrix

jacchus). J. Acoust. Soc. Am. 120, 1655–1670.

13. Bloom, K., Russell, A., and Wassenberg, K. (1987). Turn taking affects the

quality of infant vocalizations. J. Child Lang. 14, 211–227.

14. Gros-Louis, J., West, M.J., and King, A.P. (2014). Maternal responsive-

ness and the development of directed vocalizing in social interactions.

Infancy 19, 385–408.

15. Masataka, N. (1993). Effects of contingent and noncontingent maternal

stimulation on the vocal behaviour of three- to four-month-old Japanese

infants. J. Child Lang. 20, 303–312.

16. Hsu, H.-C., and Fogel, A. (2001). Infant Vocal Development in a Dynamic

Mother-Infant Communication System. Infancy 2, 87–109.

17. Goldstein, M.H., King, A.P., and West, M.J. (2003). Social interaction

shapes babbling: testing parallels between birdsong and speech. Proc.

Natl. Acad. Sci. USA 100, 8030–8035.

18. Goldstein, M.H., and Schwade, J.A. (2008). Social feedback to infants’

babbling facilitates rapid phonological learning. Psychol. Sci. 19, 515–523.

19. Margoliash, D., and Tchernichovski, O. (2015). Marmoset kids actually

listen. Science 349, 688–689.

20. Harris, R.A., Tardif, S.D., Vinar, T., Wildman, D.E., Rutherford, J.N.,

Rogers, J., Worley, K.C., and Aagaard, K.M. (2014). Evolutionary genetics

and implications of small size and twinning in callitrichine primates. Proc.

Natl. Acad. Sci. USA 111, 1467–1472.

21. Takahashi, D.Y., Fenley, A.R., and Ghazanfar, A.A. (2016). Early develop-

ment of turn-taking with parents shapes vocal acoustics in infant

marmoset monkeys. Philos. Trans. R. Soc. Lond. B Biol. Sci. 371,

20150370.

22. Tchernichovski, O., Mitra, P.P., Lints, T., and Nottebohm, F. (2001).

Dynamics of the vocal imitation process: how a zebra finch learns its

song. Science 291, 2564–2569.

23. Teramoto, Y., Takahashi, D.Y., Holmes, P., and Ghazanfar, A.A. (2017).

Vocal development in a Waddington landscape. eLife 6, e20782.

24. Fitch, W.T. (1997). Vocal tract length and formant frequency dispersion

correlate with body size in rhesus macaques. J. Acoust. Soc. Am. 102,

1213–1222.

25. Barros, M., de Souza Silva, M.A., Huston, J.P., and Tomaz, C. (2004).

Multibehavioral analysis of fear and anxiety before, during, and after exper-

imentally induced predatory stress in Callithrix penicillata. Pharmacol.

Biochem. Behav. 78, 357–367.

26. Button, K.S., Ioannidis, J.P., Mokrysz, C., Nosek, B.A., Flint, J., Robinson,

E.S., and Munafo, M.R. (2013). Power failure: why small sample size un-

dermines the reliability of neuroscience. Nat. Rev. Neurosci. 14, 365–376.

27. Gultekin, Y.B., and Hage, S.R. (2017). Limiting parental feedback disrupts

vocal development in marmoset monkeys. Nat. Commun. 8, 14046.

28. Janik, V.M., and Slater, P.J. (2000). The different roles of social learning in

vocal communication. Anim. Behav. 60, 1–11.

29. Seyfarth, R.M., and Cheney, D.L. (1986). Vocal development in vervet

monkeys. Anim. Behav. 34, 1640–1658.

30. Fischer, J. (2004). Emergence of individual recognition in young ma-

caques. Anim. Behav. 67, 655–661.

31. Zuberbuhler, K. (2000). Interspecies semantic communication in two for-

est primates. Proc. Biol. Sci. 267, 713–718.

32. Elowson, A.M., Snowdon, C.T., and Sweet, C.J. (1992). Ontogeny of trill

and J-call vocalizations in the pygmy marmoset, Cebuella pygmaea.

Anim. Behav. 43, 703–715.

33. Hihara, S., Yamada, H., Iriki, A., and Okanoya, K. (2003). Spontaneous

vocal differentiation of coo-calls for tools and food in Japanese monkeys.

Neurosci. Res. 45, 383–389.

34. Coud�e, G., Ferrari, P.F., Roda, F., Maranesi, M., Borelli, E., Veroni, V.,

Monti, F., Rozzi, S., and Fogassi, L. (2011). Neurons controlling voluntary

vocalization in the macaque ventral premotor cortex. PLoS ONE 6,

e26822.

35. Hage, S.R., Gavrilov, N., and Nieder, A. (2013). Cognitive control of distinct

vocalizations in rhesus monkeys. J. Cogn. Neurosci. 25, 1692–1701.

Current Biology 27, 1844–1852, June 19, 2017 1851

Page 10: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

36. Fitch, W.T. (2010). The Evolution of Language (Cambridge University

Press).

37. Petkov, C.I., and Jarvis, E.D. (2012). Birds, primates, and spoken language

origins: behavioral phenotypes and neurobiological substrates. Front.

Evol. Neurosci. 4, 12.

38. Tchernichovski, O., and Marcus, G. (2014). Vocal learning beyond imita-

tion: mechanisms of adaptive vocal development in songbirds and human

infants. Curr. Opin. Neurobiol. 28, 42–47.

39. Lipkind, D.,Marcus,G.F., Bemis, D.K., Sasahara, K., Jacoby, N., Takahasi,

M., Suzuki, K., Feher, O., Ravbar, P., Okanoya, K., and Tchernichovski, O.

(2013). Stepwise acquisition of vocal combinatorial capacity in songbirds

and human infants. Nature 498, 104–108.

40. West, M.J., and King, A.P. (1988). Female visual displays affect the devel-

opment of male song in the cowbird. Nature 334, 244–246.

41. Chen, Y., Matheson, L.E., and Sakata, J.T. (2016). Mechanisms underlying

the social enhancement of vocal learning in songbirds. Proceedings of the

National Academy of Sciences, 201522306.

42. Kent, R.D., and Murray, A.D. (1982). Acoustic features of infant vocalic

utterances at 3, 6, and 9 months. J. Acoust. Soc. Am. 72, 353–365.

43. Scheiner, E., Hammerschmidt, K., Jurgens, U., and Zwirner, P. (2002).

Acoustic analyses of developmental changes and emotional expression

in the preverbal vocalizations of infants. J. Voice 16, 509–529.

44. Adolph, K.E., and Robinson, S.R. (2013). The road to walking: What

learning to walk tells us about development. In Oxford Handbook of

Developmental Psychology, P. Zelazo, ed. (New York: Oxford University

Press), pp. 403–443.

45. Adolph, K.E., Vereijken, B., and Shrout, P.E. (2003). What changes in infant

walking and why. Child Dev. 74, 475–497.

46. Tamis-LeMonda, C.S., Adolph, K.E., Lobo, S.A., Karasik, L.B., Ishak, S.,

and Dimitropoulou, K.A. (2008). When infants take mothers’ advice: 18-

month-olds integrate perceptual and social information to guide motor ac-

tion. Dev. Psychol. 44, 734–746.

47. Syal, S., and Finlay, B.L. (2011). Thinking outside the cortex: social moti-

vation in the evolution and development of language. Dev. Sci. 14,

417–430.

48. Newman, S.W. (1999). The medial extended amygdala in male reproduc-

tive behavior. A node in the mammalian social behavior network. Ann. N Y

Acad. Sci. 877, 242–257.

49. Goodson, J.L. (2005). The vertebrate social behavior network: evolu-

tionary themes and variations. Horm. Behav. 48, 11–22.

50. O’Connell, L.A., and Hofmann, H.A. (2012). Evolution of a vertebrate social

decision-making network. Science 336, 1154–1157.

51. Kato, M., Okanoya, K., Koike, T., Sasaki, E., Okano, H., Watanabe, S., and

Iriki, A. (2014). Human speech- and reading-related genes display partially

overlapping expression patterns in the marmoset brain. Brain Lang. 133,

26–38.

1852 Current Biology 27, 1844–1852, June 19, 2017

52. Teramitsu, I., Kudo, L.C., London, S.E., Geschwind, D.H., and White, S.A.

(2004). Parallel FoxP1 and FoxP2 expression in songbird and human brain

predicts functional interaction. J. Neurosci. 24, 3152–3163.

53. Steiper, M.E., and Young, N.M. (2006). Primate molecular divergence

dates. Mol. Phylogenet. Evol. 41, 384–394.

54. Hrdy, S.B. (2005). Evolutionary context of human development: the

cooperative breeding model. In Attachment and Bonding: A New

Synthesis, from the 92nd Dahlem Workshop Report, C. Carter, L.

Ahnert, K. Grossmann, S. Hardy, M. Lamb, S. Porges, and N. Sachser,

eds. (Cambridge: MIT Press), pp. 9–32.

55. Burkart, J.M., Hrdy, S.B., and van Schaik, C.P. (2009). Cooperative

breeding and human cognitive evolution. Evol. Anthropol. 18, 175–186.

56. Santos, C.V., French, J.A., and Otta, E. (1997). Infant carrying behavior in

callitrichid primates: Callithrix and Leontopithecus. Int. J. Primatol. 18,

889–907.

57. Snowdon, C.T., and Cronin, K.A. (2007). Cooperative breeders do coop-

erate. Behav. Processes 76, 138–141.

58. Burkart, J.M., and van Schaik, C.P. (2010). Cognitive consequences of

cooperative breeding in primates? Anim. Cogn. 13, 1–19.

59. Yamamoto, M.E., and Lopes, F.A. (2004). Effect of Removal from the

Family Group on Feeding Behavior by Captive Callithrix jacchus. Int. J.

Primatol. 25, 489–500.

60. Burkart, J.M., Allon, O., Amici, F., Fichtel, C., Finkenwirth, C., Heschl, A.,

Huber, J., Isler, K., Kosonen, Z.K., Martins, E., et al. (2014). The evolu-

tionary origin of human hyper-cooperation. Nat. Commun. 5, 4747.

61. Ross, C., and MacLarnon, A. (2000). The evolution of non-maternal care in

anthropoid primates: a test of the hypotheses. Folia Primatol. (Basel) 71,

93–113.

62. Zuberbuhler, K. (2012). Cooperative breeding and the evolution of vocal

flexibility. In The Oxford Handbook of Language Evolution, M. Tallerman,

and K.R. Gibson, eds. (New York: Oxford University Press), pp. 71–81.

63. Locke, J.L. (2006). Parental selection of vocal behavior : Crying, cooing,

babbling, and the evolution of language. Hum. Nat. 17, 155–168.

64. Gros-Louis, J., West, M.J., Goldstein, M.H., and King, A.P. (2006).

Mothers provide differential feedback to infants’ prelinguistic sounds.

Int. J. Behav. Dev. 30, 509–516.

65. Tyler, D.E. (2008). Robust Statistics: Theory and Methods (Taylor &

Francis).

66. Hair, J.F., Anderson, R.E., Tatham, R.L., and Black, W.C. (1995).

Multivariate Data Analysis, Third Edition (New York: Macmillan).

67. Cohen, J. (1988). Statistical Power Analysis for the Behavioral Sciences

(Hillsdale, NJ: Lawrence Earlbaum Associates).

68. Olkin, I., and Finn, J.D. (1995). Correlations redux. Psychol. Bull. 118,

155–164.

69. Faul, F., Erdfelder, E., Buchner, A., and Lang, A.-G. (2009). Statistical

power analyses using G*Power 3.1: tests for correlation and regression

analyses. Behav. Res. Methods 41, 1149–1160.

Page 11: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE SOURCE IDENTIFIER

Experimental Models: Organisms/Strains

Callithrix jacchus Princeton Neuroscience Institute N/A

Software and Algorithms

MATLAB MathWorks R2015b

G*Power http://www.gpower.hhu.de/en.html v. 3.1.9.2

CONTACT FOR REAGENT AND RESOURCE SHARING

Further information and requests for resources should be directed to and will be fulfilled by the Lead Contact, Asif A. Ghazanfar

([email protected]).

EXPERIMENTAL MODEL AND SUBJECT DETAILS

MarmosetsAll experiments were approved by, and performed in compliance with, Princeton University Institutional Animal Care and Use

Committee and its guidelines. The subjects used in the study were 6 infants (3 twin sets from different parents, 3 males, 3 females,

postnatal days 1 to 60), captive commonmarmosets (Callithrix jacchus) housed at Princeton University. The number of subjects used

in this study is consistent with, or exceeds, the number of experimental animals used in other studies of nonhuman primate vocal

learning (between 1 to 4 animals) [2–4, 27]. The colony room is maintained at a temperature of approximately 27�C and 50%–

60% relative humidity, with 12L:12D light cycle. The marmosets live in family groups; all were born in captivity. They had ad libitum

access to water and were fed daily with standard commercial chow supplemented with fruits and vegetables. Additional treats were

used prior to each session to transfer the adult animals from their home-cage into a transfer cage.

METHOD DETAILS

Experimental setupStarting from their first postnatal day, we recorded the vocalizations of marmoset monkey infants in two different conditions: test

condition, in which the infants were socially isolated (i.e., undirected context) and playback condition, in which the infants received

auditory feedback from a closed loop playback system (described in more detail in the next section). Early in life, infants are always

carried by their parent. Thus, the parent carrying the infant(s) was first brought from the home cage into a transfer cage. The infant

marmoset was then gently separated from the adult and taken to the experiment room where it was placed on a flat piece of foam

inside a prism-shaped testing box made of plexiglas (0.30 m 3 0.30 m 3 0.35 m). The vocalizations we observed were identical in

type to those produced when the infant is naturally separated from parents (e.g., when parents push them off or when they transfer

them to the other parent for carrying or feeding). The experiment box rested on a table (.76m in height) in one of two opposing corners

of the room. A speaker was placed at the opposing corner, at the same height, and delivered the playback of parental calls. An opa-

que curtain made of black clothes divided the room to visually occlude the subject from the other corner. A microphone (Sennheiser

MKH-416) connected to a computer was placed directly on top of the testing box at a distance of 0.50m. Audio signals were acquired

at a sampling frequency of 96kHz at 16 bit. Most sessions consisted of a 10min test condition followed by 30min playback condition.

Between postnatal days 1 to 7, however, the test condition lasted for only 5 min to minimize the time very young infants were sepa-

rated from parents. The total number of sessions for each infant was 52, 53, 56, 56, 53, and 53 (6 infants, 317 sessions, 80,276

utterances during the test period). The number of sessions during the first 35 postnatal days were 31, 30, 33, 33, 33, 33, respectively.

The order of the infants in the twin set was counterbalanced. At each session, either the mother or father’s calls were played back.

The order of which, across sessions, the father or mother’s calls were played was counterbalanced between sessions. One of the

twin received contingent playback with low probability and the other one received contingent feedback with high probability. The

infants were randomly allocated to low or high contingency groups on first postnatal day before any measurement or procedure

was performed.

Closed loop playbacksA custommadeMATLAB routine based on Psychtoolbox-3 was used to provide playbacks in a closed loop design: it detected infant

calls and automatically delivered the parental call responses. The calls produced by infants were captured by amicrophone and sent

Current Biology 27, 1844–1852.e1–e6, June 19, 2017 e1

Page 12: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

to a desktop computer where it was saved and processed in real time. For the purpose of closed loop playback, we considered a call

any sound with an amplitude large enough to cross and stay above a fixed threshold for more than 2 s (note: immature and mature

phee calls are multi-syllabic), possibly with some silent periods each lasting less than 400 ms. Furthermore, if the ratio between the

power spectrum in the 8-10 kHz range and 4-6 kHz was larger than 2:1, then that signal was considered a more mature-sounding

contact call. When such a narrow band call was detected, a parental call was played back through a speaker at �60dB (measured

at 0.1 m from the testing box) with �1 s interval between infant call offset and playback onset. The 1 s interval was chosen based on

the distribution of parental response intervals during natural interactions (Figure S1) [21]. The parameters of the playback system

were optimized to detect infant calls and deliver playback precisely using data collected froma single infant from a separate gestation

that was not included in this study to avoid double dipping.

Detection of calls and quantification of acoustic parametersThe detection and quantification of the acoustic parameters were similar to what have been described previously [7]. To determine

the onset and offset of a syllable, a custom made MATLAB routine automatically detected the onset and offset of any signal that

differed from the background noise at specific frequency range. To detect the differences, we first bandpass filtered the entire

recording signal between 6 and 11 kHz. This corresponds to the frequency region where the infant marmoset calls have the highest

power, which is not necessarily the fundamental frequency (F0), i.e., the lowest frequency of the periodic components of the sound.

The choice of 6-11 kHz frequency range allowed us to detect 100%of calls. Second, we resampled the signal to 1 kHz sampling rate,

applied the Hilbert transform and calculated the absolute value to obtain the amplitude envelope of the signal. The amplitude enve-

lope was further low pass filtered to 50 Hz. A segment of the recording without any call (silent) was chosen as a comparison baseline.

The 99th percentile of the amplitude value in the silent period was used as the detection threshold. Sounds with amplitude envelope

higher than the threshold were considered a possible vocalization. Finally, to ensure that sounds other than vocalizations were not

included, a researcher verified whether each detected sound was a vocalization or not based on the spectrogram and amplitude of

the signal.

After detecting the onset and offset of the call syllable, a custom made MATLAB routine calculated the duration, dominant fre-

quency, amplitudemodulation (AM) frequency, andWiener entropy of each syllable. The duration of syllable is the difference between

the offset and onset of the sound amplitude that crossed the threshold. To calculate the dominant frequency of a syllable, we first

calculated the spectrogram and obtained the frequencies at which the spectrogram had maximum power for each time point.

The dominant frequency of a syllable was calculated as the maximum of those frequencies. The spectrogram was calculated using

a FFT window of 1024 points, Hanning window, with 50% overlap. The AM frequency was calculated in the following way. First, the

signal was bandpass filtered between 6 to 10 kHz and then a Hilbert transform was applied. The absolute value of the resulting signal

gives us the amplitude envelope of the modulated signal. The 6-10 kHz frequency range was found to give accurate values for the

syllable envelope. Finally, the AM frequency was calculated as the dominant frequency of the amplitude envelope. The Wiener en-

tropy is the logarithm of the ratio between the geometric and arithmetic means of the values of the power spectrum across different

frequencies [7, 22]. TheWiener entropy represents how broadband the power spectrum of a signal is. The closer the signal is to white

noise, the higher the value of Wiener entropy will be.

QUANTIFICATION AND STATISTICAL ANALYSIS

For all analysis, we adopted the Type I error a = 0.01, below which we considered statistically significant.

Multiple linear regression analysisMATLAB fitlm routine was used to fit a robust multiple linear regression to the data. The robust regression is more tolerant against

outliers, deviation from normality, heteroscedasticity in the data and is in general superior to ordinary multiple linear regression

[65]. We used the bisquare weight function with constant 4.685, which is the default in MATLAB. In Figure 2A,B we fitted the multiple

linear regression model

Entropy= a+b�PND+ c�Group+d�TwinId+ e�PND�Group+ f�PND�TwinId+ error;

where PND is the postnatal day, Group is the contingency group and TwinId is the twin identity. Both Group and TwinId were dummy

variables, which was coded as 0 or 1 for each category. Therefore, the regression model is, when written explicitly using the dummy

variables,

Entropy= a+b�PND+ c�Group+d�TwinId_1+ e�TwinId_2+ f�PND�Group+ g�PND�TwinId_1+ h�PND�TwinId_2+ error;

where, Group = 0 if it is low contingency and 1 if it is high contingency. TwinId_1 = 1 if the data is from the second twin set and

TwinId_2 = 1 if it is from the third twin set. If the data is from the first twin set, both TwinId_1 and TwinId_2 = 0. In this way all the

subject identities are coded on the dummy variables and, therefore, in our regression model, we are assuming that each marmoset

infant has its own intercept and slope. Moreover, this coding scheme allow us to control for the effect of genetics. For all other re-

gressions, Group and TwinId were always coded in the same way, but for sake of clarity, we will write the regression in the compact

form without repeating the dummy variables (Group and TwinId).

e2 Current Biology 27, 1844–1852.e1–e6, June 19, 2017

Page 13: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

The multiple linear model fitted in Figures 2C and 2D was:

AMfreq= a+b�PND+ c�Group+d�TwinId+ e�PND�Group+ f�PND�TwinId+ error

For the regression controlling for weights, we fitted the multiple linear models:

Entropy= a+b�PND+ c�Group+d�TwinId+ e�Weight+ f�PND�Group+g�PND�TwinId+ + h� PND�Weight+ error

� � � �

AMFreq= a+b PND+ c Group+d TwinId+ e Weight+ f�PND�Group+ g�PND�TwinId+ + h� PND�Weight+ error

The cubic spline curve was fitted to the weights of each infant usingMATLAB fitlm. Anymissing weight data was interpolated using

the values of the cubic spline curve.

For the multiple linear regressions in Figures 3, the models had exactly the same predictors as in the models for Figure 2, but using

as dependent variables the duration of syllables and dominant frequency.

To test the effect of total number of stimuli calls on the previous session, we also fitted the multiple linear regressions

Dependent= a+b�PND+ c�Group+d�TwinId+ e� NStim+ f�PND�Group+g�PND�TwinId+ h�PND�NStim+ error;

where the dependent variable was either entropy, AM frequency, duration or dominant frequency and NStim is the number of feed-

back stimuli that the infant received in the previous session. Because, the interaction term between PND and NStim was not statis-

tically significant for any of the acoustic parameters, we excluded the interaction term and reduced the model to

Dependent= a+b�PND+ c�Group+d�TwinId+ e� NStim+ f�PND�Group+g�PND�TwinId+ error:

By experimental design, NStim is correlated with Group. To verify if the degree of collinearity of these predictors could cause dif-

ficulty in the interpretation of the coefficients, we calculated the variance inflation factor (VIF). Values of VIF larger than 10 are usually

considered problematic [66]. We found that the VIF of NStim is 4.7.

The linear regression fitted to the data shown in Figure 4 was

CallRate= a+b�PND+ c�Group+d�TwinId+ e�PND�Group+ f�PND�TwinId+ error;

where CallRate is the rate of call production by the infant in test condition.

Linear regression modelsWe report below the estimated regression coefficients, standard errors, t-values, and p values of themodels used to test the effect of

interaction between postnatal day and contingency group. All models were tested against the constant model and were significantly

different (p < 0.0001).

Entropy �1 + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

__________ ________ ________ __________

(Intercept) �1.0732 0.2167 �4.9523 1.6461e-06

PND �0.028009 0.010559 �2.6527 0.0086788

Group �0.058207 0.21319 �0.27302 0.78514

TwinId_1 �0.73452 0.26191 �2.8044 0.005579

TwinId_2 �0.21587 0.26217 �0.8234 0.41134

PND:Group �0.032279 0.01041 �3.1009 0.0022313

PND:TwinId_1 �0.0013244 0.012797 �0.10349 0.91768

PND:TwinId_2 0.049178 0.012803 3.8411 0.00016818

AMfreq �1 + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

________ ______ ________ __________

(Intercept) 1107.7 59.223 18.703 1.6341e-44

PND �0.3784 2.8856 �0.13113 0.89581

Group 105.62 58.264 1.8129 0.071474

TwinId_1 �444.54 71.578 �6.2106 3.3977e-09

TwinId_2 152.61 71.648 2.1299 0.034499

PND:Group �7.7853 2.8448 �2.7366 0.0068124

PND:TwinId_1 3.3999 3.4974 0.97213 0.33226

PND:TwinId_2 �10.507 3.499 �3.003 0.0030433

Current Biology 27, 1844–1852.e1–e6, June 19, 2017 e3

Page 14: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

e4

Duration �1 + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

__________ _________ _______ __________

(Intercept) 0.2592 0.02994 8.6573 2.3366e-15

PND 0.009027 0.0014588 6.188 3.8273e-09

Group 0.036144 0.029455 1.2271 0.22135

TwinId_1 0.10384 0.036186 2.8696 0.0045888

TwinId_2 0.042662 0.036222 1.1778 0.24039

PND:Group 0.0028045 0.0014382 1.95 0.052687

PND:TwinId_1 �0.0040556 0.0017681 �2.2938 0.022925

PND:TwinId_2 �0.0086818 0.0017689 �4.908 2.0113e-06

DomFreq �1 + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

________ ______ _________ __________

(Intercept) 10234 94.532 108.26 3.018e-169

PND �10.81 4.606 �2.3468 0.019992

Group �174.27 93.001 �1.8738 0.062531

TwinId_1 �450.33 114.25 �3.9415 0.00011473

TwinId_2 �3.813 114.37 �0.033341 0.97344

PND:Group 2.2053 4.541 0.48566 0.62779

PND:TwinId_1 �14.693 5.5825 �2.6321 0.0092037

PND:TwinId_2 �8.08 5.5851 �1.4467 0.14967

Entropy �1 + PND*Group + PND*TwinId + PND*Weight

Estimated Coefficients:

Estimate SE tStat pValue

_________ __________ _________ __________

(Intercept) 1.0381 1.2497 0.83068 0.40724

PND �0.065573 0.026399 �2.4839 0.013895

Group 0.19384 0.26076 0.74334 0.45823

TwinId_1 �0.22484 0.39588 �0.56794 0.57077

TwinId_2 �0.016016 0.28399 �0.056397 0.95509

Weight �0.079176 0.049991 �1.5838 0.11497

PND:Group �0.03973 0.010633 �3.7367 0.00024906

PND:TwinId_1 0.0078016 0.026358 0.29598 0.76758

PND:TwinId_2 0.046426 0.01376 3.3741 0.0009041

PND:Weight 0.0019526 0.00069189 2.8221 0.0052986

AMfreq �1 + PND*Group + PND*TwinId + PND*Weight

Estimated Coefficients:

Estimate SE tStat pValue

________ _______ _______ __________

(Intercept) 2509.2 309.91 8.0965 7.7001e-14

PND 30.551 6.5469 4.6665 5.9033e-06

Group 277.62 64.668 4.2931 2.8539e-05

TwinId_1 �121.99 98.176 �1.2426 0.21561

TwinId_2 284.48 70.428 4.0394 7.8784e-05

Weight �60.478 12.398 �4.8782 2.318e-06

PND:Group �7.3904 2.6368 �2.8028 0.0056124

PND:TwinId_1 34.435 6.5366 5.2679 3.8493e-07

PND:TwinId_2 �5.6191 3.4123 �1.6467 0.10134

PND:Weight 0.29006 0.17159 1.6905 0.092639

Current Biology 27, 1844–1852.e1–e6, June 19, 2017

Page 15: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

Duration �1 + PND*Group + PND*TwinId + PND*Weight

Estimated Coefficients:

Estimate SE tStat pValue

__________ __________ _______ __________

(Intercept) 0.086621 0.17803 0.48656 0.62715

PND 0.0034503 0.003761 0.91741 0.36013

Group 0.026253 0.037149 0.70669 0.48066

TwinId_1 0.080725 0.056398 1.4313 0.15404

TwinId_2 0.024033 0.040458 0.59403 0.55323

Weight 0.0072414 0.0071219 1.0168 0.3106

PND:Group 0.0016772 0.0015147 1.1072 0.26965

PND:TwinId_1 �0.0093137 0.003755 �2.4803 0.01403

PND:TwinId_2 �0.0084514 0.0019603 �4.3114 2.6477e-05

PND:Weight 1.4905e-05 9.8569e-05 0.15121 0.87998

DomFreq �1 + PND*Group + PND*TwinId + PND*Weight

Estimated Coefficients:

Estimate SE tStat pValue

________ _______ _________ _________

(Intercept) 10624 560.78 18.946 5.199e-45

PND �0.4841 11.847 �0.040863 0.96745

Group �160.03 117.02 �1.3676 0.17312

TwinId_1 �356.62 177.65 �2.0074 0.046174

TwinId_2 56.528 127.44 0.44356 0.65788

Weight �16.719 22.434 �0.74528 0.45706

PND:Group 4.951 4.7714 1.0376 0.30081

PND:TwinId_1 �5.6357 11.828 �0.47646 0.63431

PND:TwinId_2 �8.7727 6.1747 �1.4208 0.15709

PND:Weight 0.0261 0.31049 0.084062 0.9331

Entropy �1 + NStim + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

_________ _________ ________ __________

(Intercept) �1.0594 0.21981 �4.8195 3.0015e-06

PND �0.028755 0.010712 �2.6844 0.0079286

Group 0.12121 0.2956 0.41006 0.68224

TwinId_1 �0.75217 0.26418 �2.8472 0.0049126

TwinId_2 �0.2013 0.26559 �0.75793 0.44946

NStim �0.00163 0.0018645 �0.87424 0.38313

PND:Group �0.031149 0.010617 �2.9338 0.0037743

PND:TwinId_1 0.0021952 0.012975 0.16918 0.86584

PND:TwinId_2 0.050836 0.013021 3.9042 0.00013259

AMfreq �1 + NStim + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

________ _______ _______ __________

(Intercept) 1065.3 58.295 18.274 3.2489e-43

PND 1.5478 2.8408 0.54485 0.58652

Group �95.55 78.395 �1.2188 0.22447

TwinId_1 �411.68 70.062 �5.8759 1.9362e-08

TwinId_2 168.22 70.437 2.3883 0.017939

NStim 1.665 0.49446 3.3673 0.00092433

PND:Group �8.1971 2.8158 �2.9112 0.0040459

PND:TwinId_1 0.58617 3.441 0.17035 0.86492

PND:TwinId_2 �13.907 3.4532 �4.0273 8.244e-05

Current Biology 27, 1844–1852.e1–e6, June 19, 2017 e5

Page 16: Vocal Learning via Social Reinforcement by Infant Marmoset Monkeysdtakahas/publications/TakahashiLiaoGh... · 2019. 8. 1. · Current Biology Report Vocal Learning via Social Reinforcement

e6

Duration �1 + NStim + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

__________ __________ _______ __________

(Intercept) 0.25613 0.030179 8.4869 6.9102e-15

PND 0.0091292 0.0014707 6.2074 3.4868e-09

Group 0.0062594 0.040585 0.15423 0.8776

TwinId_1 0.10452 0.036271 2.8816 0.0044266

TwinId_2 0.044412 0.036465 1.2179 0.22481

NStim 0.00024752 0.00025598 0.96693 0.33485

PND:Group 0.0028697 0.0014577 1.9686 0.050497

PND:TwinId_1 �0.0043113 0.0017814 �2.4201 0.016487

PND:TwinId_2 �0.0093599 0.0017877 �5.2356 4.461e-07

DomFreq �1 + NStim + PND*Group + PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

________ _______ ________ ___________

(Intercept) 10221 98.247 104.03 2.0214e-165

PND �10.317 4.7878 �2.1548 0.032478

Group �424.87 132.12 �3.2157 0.0015373

TwinId_1 �409.35 118.08 �3.4667 0.00065583

TwinId_2 6.0924 118.71 0.051321 0.95913

NStim 1.6855 0.83334 2.0226 0.04456

PND:Group 4.6416 4.7456 0.97809 0.32931

PND:TwinId_1 �18.061 5.7994 �3.1142 0.0021395

PND:TwinId_2 �12.989 5.8199 �2.2318 0.026832

CallRate�PND*Group+PND*TwinId

Estimated Coefficients:

Estimate SE tStat pValue

________ ________ _______ __________

(Intercept) 16.342 1.9319 8.4591 8.0002e-15

PND �0.26321 0.09413 �2.7962 0.0057162

Group_1 �2.286 1.9006 �1.2028 0.23061

TwinId_2 �5.2229 2.3349 �2.2369 0.02649

TwinId_3 9.6716 2.3372 4.1381 5.3129e-05

PND:Group_1 0.040069 0.092801 0.43177 0.66641

PND:TwinId_2 0.05292 0.11409 0.46386 0.6433

PND:TwinId_3 0.1245 0.11414 1.0908 0.27678

Effect size and power analysisWe calculated the local effect size of the contingency group (Group) for the model in Figure 2 and Figure 3. We used as a measure of

the effect size the Cohen’s f2 [67]. To calculate the confidence interval, we used the Olkin and Finn’s approximation [68].The power

was calculated using G*Power 3 [69].

DATA AND SOFTWARE AVAILABILITY

Data andMATLAB code used for analysis of Figures 2, 3, and 4 are available in DRYAD Digital Repository: http://dx.doi.org/10.5061/

dryad.76bn8.

Current Biology 27, 1844–1852.e1–e6, June 19, 2017